Regio- and Stereodivergent Allylic Reductions of Bicyclic Piperidine

7 hours ago - User Resources. About Us · ACS Members · Librarians · ACS Publishing Center · Website Demos · Privacy Policy · Mobile Site ...
0 downloads 0 Views 698KB Size
Subscriber access provided by Kaohsiung Medical University

Note

Regio- and Stereodivergent Allylic Reductions of Bicyclic Piperidine Enecarbamate Derivatives Francesco Berti, Andrea Menichetti, Lucilla Favero, Fabio Marchetti, and Mauro Pineschi J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.8b01601 • Publication Date (Web): 13 Sep 2018 Downloaded from http://pubs.acs.org on September 13, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Regio- and Stereodivergent Allylic Reductions of Bicyclic Piperidine Enecarbamate Derivatives Francesco Berti,§ Andrea Menichetti, § Lucilla Favero, § Fabio Marchetti,† and Mauro Pineschi§* §

Dipartimento di Farmacia, Sede di Chimica Bioorganica e Biofarmacia, Università di Pisa, Via Bonanno 33, 56126 Pisa, Italy.



Dipartimento di Chimica e Chimica Industriale, Università di Pisa, via G. Moruzzi 3, 56124,

Pisa, Italy. RECEIVED DATE (to be automatically inserted after your manuscript is accepted if required according to the journal that you are submitting your paper to)

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

ABSTRACT: The particular nature of tetrahydropyrido[4,3-e]-1,4,2-dioxazines of type 1 allows

the

regio-

and

stereoselective

obtainment

of

substituted

N-carbamoyl

tetrahydropyridines by common reducing agents. A completely novel, biologically active, bicyclic 1,3-diaza-4-oxa-[3.3.1]-nonene scaffold can be generated by the use of lithium triethylborohydride through an unprecedented cascade syn-SN2’ reduction/carbamate reduction/cyclization reactions. The remarkable regioselectivity switches of allylic reduction process have been rationalized with the aid of computational studies.

Tetrahydropyridines are important targets in synthetic organic chemistry due to their relevance as privilege structure in medicinal chemistry.1 Therefore, considerable efforts have been devoted to the individuation of new methods for their regio- and stereoselective synthesis and functionalization.2 However, a regio- and stereodivergent reduction of piperidine enecarbamates bearing an allylic leaving group has not yet been reported. The main reason for this shortcoming can be found in the relative instability of 1,2,3,4tetrahydropyridine derivatives containing common leaving groups at the C4-position. We recently became interested in the elaboration of nitroso cycloadducts derived from 2substituted-1,2-dihydropyridines,3 easily accessible by addition of nucleophiles to activated pyridinium salts (eq. a, Scheme 1).4,5 The [3,3]-hetero Cope rearrangement of the obtained nitroso Diels-Alder (NDA) cycloadducts has been reported to give 4a,7,8,8atetrahydropyrido[4,3-e]-1,4,2-dioxazines of type 1 (eq. a, Scheme 1).6 We noticed that the study of their reactivity is limited to the N-O reductive cleavage with Raney nickel to obtain racemic aminoarabinose and aminoaltrose derivatives.6a On the other hand, palladiumcatalyzed allylic reduction have been used extensively in synthetic organic chemistry to

ACS Paragon Plus Environment

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

introduce hydride species into a variety of allylic substrates in a regio- and stereoselective fashion.7-9 We herein report a new method to regio- and stereoselectively reduce piperidine enecarbamates, with the individuation of a fused 1,2,4-dioxazine framework as an unconventional allylic leaving group (see compound of type 1, Scheme 1). At the outset of this study, we identified optimal reaction conditions to obtain 1,2,4dioxazine derivatives of type 1 reproducibly and in satisfactory yields (eq. b, Scheme 1).10 While attempting to reduce the N-O bond of racemic 1,4,2-dioxazine 1a, easily prepared in gram scale and taken as a model substrate, we serendipitously found that a heterogeneous hydrogenation carried out with ammonium formate at 80 °C in the presence of Pd/C (20 mol%) at 80 °C readily afforded a 60/40 mixture of saturated derivatives 2a and 3a (eq. c, Scheme 1).11

Scheme 1. State of the art, synthesis and preliminary reduction of tetrahydropyrido[4,3-e]1,4,2-dioxazines (1).

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 30

A result of this kind can be reasonably explained by admitting the intervention of 4,5dehydro-intermediate 4a by initial hydride transfer at the position adjacent to the nitrogen, followed by an allylic deoxygenation of the C3-O bond. Stimulated by these preliminary results, we decided to carry out a deeper investigation of the reduction of enecarbamate 1a with hydride species in order to develop selective allylic reduction processes.8b,12 While sodium borohydride was totally ineffective, the use of the

lithium

analog

(LiBH4)

afforded

2,3-dehydropiperidine

5a

with

complete

regioselectivity in attack of the hydride at the C4 position (Scheme 2). It should be noted that 3-oxygenated-2-aryl-piperidines such as 5a are compounds of considerable importance in medicinal chemistry.13

Scheme 2. Reduction of 1,2,4-dioxazine 1a with metal hydride species. Much to our surprise, the reaction of 1a with an excess of highly reactive lithium triethylborohydride (Super-H®) afforded readily a good yield of compound 6a possessing a novel 1,3-diaza-4-oxa-[3.3.1]-bicyclic scaffold, whose structure was unequivocally

ACS Paragon Plus Environment

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

determined by single crystal X-ray analysis (Figure 1).14 This reaction to can be easily scaled-up to 1.0 mmol scale with only marginal deviations from the initial reaction conditions carried out on a 0.15 mmol scale. The use of analog sodium salt (NaEt3BH) afforded the same compound with a lower yield and an increased reaction time. The use of LiAlH4 afforded a complex mixture of products in which only 6a could be detected.

Figure 1. ORTEP diagram of 6a. Thermal ellipsoids are at 30% probability. Relevant bond distances (Å): C1−N1 1.468(2), C1−C2 1.498(2), C2−C3 1.315(2), C3−C4 1.490(2), C4−O1 1.483(1), C4−C5 1.524(1), C5−N1 1.479(1), N1−C12 1.454(1), O1−N2 1.403(1), N2−C12 1.460(1).

In order to get more insight into the reaction mechanism we carried out the reaction using LiEt3BD as the reducing agent (Figure 2).

Figure 2. Plausible mechanism for the formation of the 1,3-diaza-4-oxa-[3.3.1]-nonene framework.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

NMR analysis of deuterated bicyclic compound 6a-D showed clearly the regio- and stereochemistry of deuterium incorporation.15 The only way to obtain 1,3-diaza-4-oxa[3.3.1]-bicyclic framework 6a comes from a regioselective SN2’-addition of the hydride to enecarbamate 1a, which turned out to be syn to the dioxazine leaving group. The synstereoselectivity can be reasonably ascribed to simultaneous coordination of the cation to the oxygen of the leaving group and the hydride of the M-HBEt3 on conformation 1a-B (see Supporting Information).16,17 The obtained anionic 3,4-unsaturated intermediate A does not undergo a further reduction of the oxyamido functionality arguably due to a intramolecular chelate stabilization. It is also likely that the concomitant reduction of the carbamate moiety of species A gives rise to zwitterionic iminium ion B that undergoes intramolecular trapping by N-alkylation to deliver the bicyclic framework 6a-D (Figure 2). In accordance with the proposed mechanism, the formation of such a bicyclic framework necessarily requires the presence of a carbamate protecting group on the endocyclic piperidine nitrogen. As a matter of fact, the LiEt3BH reduction of enamides 1b-d afforded a complex mixture of products not containing the corresponding oxadiazabicyclononene scaffold (Scheme 3). Hence, only using enecarbamates 1e-1k it was possible to obtain the corresponding bicyclic structures 6b-g, albeit with variable yields (Scheme 3).

ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Scheme 3. Scope of the formation of [3.3.1]-bicyclic scaffold.

The first step (SN2’-reduction) of this sequential reaction is consistent with the known ability of highly nucleophilic and soft reagent lithium triethylborohydride to undergo addition to double bonds of different nature including the fast reduction of pyridine to tetrahydropyridine.18 NBO calculations showed that π*C=C (+0.0269 au) is less energetic than σ*C-O(+0.2330 au) and therefore a softer reagent should have a preference for the C2 position. On the other hand, by examination of an electrostatic map of the fused dioxazine 1a it is clear that the C4 is more electropositive than C2 carbon (Figure 3). Evidently, the C4 is the more reactive position when dealing with harder LiBH4, a reagent that is known to favor 1,2-addition products in conjugated systems.19

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. Molecular electrostatic potential of dioxazine 1a.

Having established that the C4-oxygen part of the 1,2,4-dioxazine framework can act as a leaving group in regioselective allylic displacement reactions with metal hydrides, we were intrigued how palladium catalysis could influence such reactivity on model compound 1a. The electrophilic Pd(II)-center activates the allyl system for the nucleophilic attack at the allylic termini, which is usually the rate limiting step.8 It also commonly accepted that hydride as nucleophile is initially transferred to palladium and from there to the allylic ligand by reductive elimination.8,9 When 1a was allowed to react with sodium borohydride in the presence of catalytic amounts of (PPh3)2PdCl2 a sluggish reaction now occurred to afford an equimolar mixture of compounds 5a (SN2-reduction) and 4a (SN2’-reduction) (Table 1, entry 1). The same reaction carried out with NaBD4 afforded compound 4a-D with the deuterium atom incorporated in a trans fashion to the 2-phenyl group (entry 2).15 This clearly indicated that the syn-SN2’-reduction pathway is external to the coordination sphere of palladium. On the other hand, in this reaction compound 5a-D was obtained with

ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

inversion of configuration with respect to the leaving group (i.e. from the same side of palladium).15 When the reaction of 1a with sodium borohydride was carried out in the presence of catalytic amounts of (CH3CN)2PdCl2, compound 5a was obtained with a complete regiocontrol (entry 3).

Table 1. Reduction of 1a with metal hydrides in the presence of palladium salts.

entrya

M-H

Pd

t[h]

4a/5a/6ab

Yield [%]c

1

NaBH4

(PPh3)2PdCl2

24

60/40/0

45

2

NaBD4

(PPh3)2PdCl2

8

43/57/0

37

3

NaBH4

(CH3CN)2PdCl2

32

0/100/0

25

4

LiBH4

(PPh3)2PdCl2

1

28/72/0

50

5

LiBH4

(CH3CN)2PdCl2

1

0/100/0

58

6

LiAlH4

(PPh3)2PdCl2

1

7

LiEt3BH

(PPh3)2PdCl2

0.5

0/10/90

65

8

LiEt3BH

[(PPh3)]4Pd

0.5

0/3/97

75

9

LiEt3BH

Pd(dba)2

1

0/63/37

53

10

LiEt3BH

(CH3CN)2PdCl2

0.5

0/100/0

60

11

NaEt3BH

(CH3CN)2PdCl2

1

0/100/0

55

complex mixture

a

Unless stated otherwise, all reactions were carried out at 0 °C using 1a (0.15 mmol), Pd-salt (0.0075 mmol), THF (0.75 mL), metal hydride (0.9 mmol).b Determined by 1H NMR of the crude mixture (value of 0 means not detected by NMR). c Isolated yields of products after chromatographic purification on SiO2.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The examination of other hydrides per se able to reduce 1a and therefore in competition with the palladium-catalyzed reduction was rather instructive. Palladium-catalysis gave a clear acceleration of the reduction process using lithium borohydride, but only (PPh3)2PdCl2 modified the inherent regioselectivity of the reaction with lithium borohydride (entries 4 and 5). The inherent reactivity of Super-hydride® with 1a was only marginally affected by the presence of (PPh3)2PdCl2 and Pd[(PPh3)]4 (entries 7 and 8). The use of Pd(dba)2 gave a mixture of regioisomers with a prevalence of SN2-addition product 5a (entry 9). Interestingly, a complete reversal of regioselectivity with the exclusive formation of 1,2,3,4-tetrahydropyridine 5a was obtained using LiEt3BH or NaEt3BH in the presence of catalytic amount of (CH3CN)2PdCl2 (entries 10 and 11). Hence, it is plausible that when only more coordinating ligands are used, the inherent reactivity of the hydride species can compete with the reductive elimination pathway from palladium. A computational study performed at the DFT level revealed the unsymmetrical structure of the corresponding PdL2-allyl complex with a Pd-C4 bond strongly shorter of the Pd-C2 bond (Figure 4), regardless of the palladium ligand, with a consequent marked carbocationic character of C(2) atoms, as confirmed by NBO analysis (+0.188 au for L=CH3CN and +0.184 au for L=Ph3P).17

ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Figure 4. DFT optimized structure and NBO Charges (Pd and C(2) atoms) of Pd-allyl complex derived from (CH3CN)2PdCl2

At the same time Pd atom is more electropositive in the Pd-allyl complex derived from (CH3CN)2PdCl2 (NBO Charge +0.282 au) than in the one derived from (Ph3P)2PdCl2 (NBO Charge +0.036 au). It is plausible that a more positive palladium (L=CH3CN) undergoes more readily the reduction and the subsequent reductive elimination at the C4-position, whereas in the case of a less positive palladium (L=Ph3P) reduction at C(2) becomes competitive. In summary, we have developed new regioselective reductions of piperidine enecarbamates bearing a stable allylic leaving group. By the use of common reducing agents, with or without palladium catalysts, it is possible to obtain selectively ∆2,3- or ∆3,4piperidine derivatives with unconventional substitution patterns. By transient formation of the latter derivative, a new stable oxadiaza-[3.3.1]-bicyclic scaffold can be obtained by a robust cascade reaction. To our delight, compound 6a selectively stimulated GLP-1 secretion in mouse enteroendocrine-like STC-1 cells with an EC50 = 11 M. Further studies

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 30

are ongoing to assess the biological activity of a library of compounds based on this original scaffold.

EXPERIMENTAL SECTION General Information. All reactions dealing with air or moisture sensitive compounds were carried out under an argon atmosphere in oven dried 10 mL or 25 mL Schlenk tubes. Analytical TLC were performed on Merck TLC Silica gel 60 silica gel sheets with detection by exposure to ultraviolet light (254 nm) and/or by immersion in an acidic staining solution of p-anisaldehyde in EtOH. Silica gel 60 was used for flash chromatography (230400 mesh). Automated column chromatography was performed using prepacked silica gel cartridges on a Biotage Isolera 1.5.2 (27-53 μm). 1H NMR spectra were recorded on Bruker Avance II 250 MHz spectrometer or on Bruker Avance III 300 MHz spectrometer. Chemical shifts are reported in ppm downfield from tetramethylsilane with the solvent resonance as the internal standard (deuterochloroform: δ 7.26, deuteroacetonitrile: δ 1.94).

13C

NMR

spectra were recorded on a Bruker Avance II 250 spectrometer (62.5 MHz) with complete proton decoupling. Chemical shifts are reported in ppm downfield from tetramethylsilane with the solvent resonance as the internal standard (deuterochloroform: δ 77.16, deuteroacetonitrile: δ 1.32). Melting points were determined on a Kofler apparatus and are uncorrected. Mass spectra ESIMS were measured on a Finnigan LC-Q Deca Termoquest spectrometer, equipped with a software Xcalibur. Solvents for extraction and chromatography were distilled before use. Compounds of type 1 were prepared in accordance to the indicated literature procedures.10 LiBDEt3 was prepared by procedure of Brown.20

ACS Paragon Plus Environment

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

General procedure for the synthesis of inverse cycloadducts A1-3 (Scheme S1, Supporting Information): a round-bottomed flask was charged with the appropriate 1,2dihydropyridine and N-hydroxy-2-phenylacetamide in a 3:1 solution of MeOH/H2O (0.17 M) at 0 °C. NaIO4 (1.1 eq, 99.5% purity) was slowly added and the resulting heterogeneous mixture was allowed to stir until no 1,2-dihydropyridine was detected by TLC analysis. The reaction crude was concentrated and diluted with CH2Cl2/H2O to give a biphasic solution. The aqueous layer was extracted with CH2Cl2 and the combined organic layers were dried over MgSO4, filtered and concentrated. The resulting residue was subjected to flash chromatography.

Methyl

(1S*,4R*)-3-(2-phenylacetyl)-6-vinyl-2-oxa-3,5-diazabicyclo[2.2.2]oct-7-ene-5-

carboxylate (A1). According to the general procedure, methyl 2-vinylpyridine-1(2H)carboxylate (300 mg, 1.8 mmol),21 N-hydroxy-2-phenylacetamide (302 mg, 2.0 mmol), NaIO4 (430 mg, 2.0 mmol), MeOH (7.2 mL) and H2O (3.6 mL). Subsequent flash chromatography (petroleum ether/AcOEt 7:3 + 2% Et3N, Rf = 0.33) afforded the title compound as a colorless oil (216 mg, 38%). 1H NMR (250 MHz, CD3CN, 60 °C) δ 7.37 – 7.17 (m, 5H), 6.80 – 6.65 (m, 2H), 6.44 – 6.34 (m, 1H), 5.61 – 5.43 (m, 1H), 5.22 (dt, J = 11.2, 1.4 Hz, 1H), 5.17 (dt, J = 4.2, 1.4 Hz, 1H), 4.93 – 4.88 (m, 1H), 4.57 – 4.50 (m, 1H), 3.69 (s, 3H), 3.66 – 3.55 (m, 2H). 13C NMR (62.5 MHz, CD3CN) δ 174.5, 156.7, 135.3, 134.8, 132.9, 130.6, 129.3, 129.2, 127.6, 118.6, 74.9, 60.0, 59.3, 53.5, 40.5. HRMS (ESI) m/z [M + Na+] Calcd for C17H18N2O4Na 337,1164, found 337,1159.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Methyl

(1S*,4R*)-6-cyclohexyl-3-(2-phenylacetyl)-2-oxa-3,5-diazabicyclo[2.2.2]oct-7-

ene-5-carboxylate (A2). According to the general procedure, methyl 2-cyclohexylpyridine1(2H)-carboxylate (1.4 g, 2.5 mmol, 40 % purity),21 N-hydroxy-2-phenylacetamide (422 mg, 2.8 mmol), NaIO4 (599 mg, 2.8 mmol), MeOH (10 mL) and H2O (5 mL). Subsequent flash chromatography (petroleum ether/AcOEt 7:3 + 2% Et3N, Rf = 0.18) afforded the title compound as a yellow oil (444 mg, 48%). 1H NMR (250 MHz, CD3CN) δ 7.37 – 7.16 (m, 5H), 6.74 – 6.65 (m, 1H), 6.65 – 6.57 (m, 1H), 6.51 – 6.41 (m, 1H), 5.03 – 4.95 (m, 1H), 3.90 (dd, J = 6.8, 3.6 Hz, 1H), 3.82 – 3.78 (m, 1H), 3.71 (s, 3H), 3.58 (ABq, J = 15.1 Hz, 2H), 1.73 (s, 2H), 1.68 – 1.43 (m, 4H), 1.28 – 1.01 (m, 5H). 13C NMR (62.5 MHz, CD3CN) δ 175.8, 158.5, 136.0, 133.0, 131.3, 130.9, 129.9, 128.2, 74.7, 64.0, 61.3, 54.3, 41.5, 41.3, 31.3, 31.2, 27.8, 27.7, 27.7. HRMS (ESI) m/z [M + Na+] Calcd for C21H26N2O4Na 393,1790, found 393,1786.

Methyl

(1S*,4R*)-6-allyl-3-(2-phenylacetyl)-2-oxa-3,5-diazabicyclo[2.2.2]oct-7-ene-5-

carboxylate (A3). According to the general procedure, methyl 2-allylpyridine-1(2H)carboxylate (716 mg, 4.0 mmol),22 N-hydroxy-2-phenylacetamide (665 mg, 4.4 mmol), NaIO4 (940 mg, 4.4 mmol), MeOH (16 mL) and H2O (8 mL). Subsequent flash chromatography (petroleum ether/Et2O 6:4 + 2% Et3N, Rf = 0.18) afforded the title compound as a yellow oil (603 mg, 46%). 1H NMR (250 MHz, CD3CN, 60 °C) δ 7.38 – 7.15 (m, 5H), 6.79 – 6.69 (m, 1H), 6.67 – 6.58 (m, 1H), 6.50 – 6.39 (m, 1H), 5.90 – 5.69 (m, 1H), 5.13 (dd, J = 4.7, 3.4 Hz, 1H), 5.07 (bs, 1H), 4.96 – 4.86 (m, 1H), 4.07 – 3.95 (m, 1H), 3.60 (ABq, J = 15.3 Hz, 2H), 3.72 (s, 3H), 2.61 – 2.47 (m, 1H), 2.03 – 1.84 (m, 1H). 13C NMR (62.5 MHz, CD3CN), major rotamer δ 174.6, 156.5, 136.0, 135.3, 134.5, 132.9, 130.6, 129.2, 127.6,

ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

118.4, 74.3, 59.6, 58.0, 53.6, 40.6, 36.8. HRMS (ESI) m/z [M + Na+] Calcd for C18H20N2O4Na 351,1321, found 351.1317.

General procedure for the synthesis of the dioxazine derivatives 1i-k:10 an oven-dried 10 mL pyrex vial was charged with the specified inverse cycloadduct of type A, copper(I) chloride (0.20 eq) and 1,2-dichloroethane (DCE) (0.15 M). The resulting mixture was allowed to react until no compound A1-3 was detected by TLC analysis. The reaction was quenched with H2O and the aqueous phase was extracted with CH2Cl2 and the combined organic layers were dried over MgSO4. Removal of the solvent gave a residue that was purified by flash chromatography.

(4aR*,8S*,8aS*)-Methyl

3-benzyl-8-vinyl-8,8a-dihydropyrido[4,3-e][1,4,2]dioxazine-

7(4aH)-carboxylate (1i). According to the general procedure, A1 (193 mg, 0.61 mmol), CuCl (12.2 mg, 0.123 mmol) in DCE (4.10 mL) reacted at 75 °C for 17 h. Subsequent flash chromatography (petroleum ether /AcOEt 7:3, Rf = 0.23) afforded the title compound as a colorless oil (140 mg, 72.5 %). 1H NMR (250 MHz, CD3CN, 65 °C) δ 7.39 – 7.18 (m, 5H), 6.91 (d, J = 8.5 Hz, 1H), 5.74 (ddd, J = 15.7, 10.6, 4.8 Hz, 1H), 5.26 (d, J = 10.5 Hz, 1H), 5.18 (d, J = 17.4 Hz, 1H), 4.94 (bs, 1H), 4.79 (bs, 1H), 4.66 (d, J = 8.3 Hz, 1H), 4.06 – 3.98 (m, 1H), 3.75 (s, 3H) 3.48 (s, 2H). 13C NMR (62.5 MHz, CD3CN) δ 156.5, 154.6, 136.9, 131.5, 129.6, 129.5, 127.9, 127.1, 117.9, 102.1, 67.1, 66.7, 57.3, 54.1, 38.6. HRMS (ESI) m/z [M + Na+] Calcd for C17H18N2O4Na 337,1164, found 337.1159.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4aR*,8S*,8aS*)-Methyl

3-benzyl-8-cyclohexyl-8,8a-dihydropyrido[4,3-

e][1,4,2]dioxazine-7(4aH)-carboxylate (1j). According to the general procedure, A2 (69 mg, 0.18 mmol), CuCl (3.7 mg, 0.037 mmol) in DCE (1.3 mL) reacted at 75 °C for 2 h. Subsequent flash chromatography (petroleum ether /AcOEt 8:2, Rf = 0.25) afforded the title compound as a white amorphous solid (40 mg, 58 %). 1H NMR (250 MHz, CD3CN, 65 °C) δ 7.39 – 7.20 (m, 5H), 6.83 (d, J = 7.7 Hz, 1H), 4.92 (bs, 1H), 4.70 (d, J = 7.5 Hz, 1H), 4.19 (d, J = 8.9 Hz, 1H), 4.07 (s, 1H), 3.73 (s, 1H), 3.47 (s, 2H), 1.82 – 1.40 (m, 6H), 1.32 – 0.98 (m, 5H).13C NMR (62.5 MHz, CD3CN) δ 156.6, 155.3 and 155.2*, 136.9, 129.6, 129.5, 127.8, 127.6 and 127.3*, 103.2* and102.7, 67.5, 65.5, 60.0, 54.0 and 53.7*, 38.6, 37.9, 30.3, 26.8, 26.6, 26.5. [* minor rotamer]. HRMS (ESI) m/z [M + Na+] Calcd for C21H26N2O4Na 393,1790, found 3931784.

(4aR*,8S*,8aS*)-Methyl

3-benzyl-8-allyl-8,8a-dihydropyrido[4,3-e][1,4,2]dioxazine-

7(4aH)-carboxylate (1k). According to the general procedure, A3 (70 mg, 0.21 mmol), CuCl (4.2 mg, 0.043 mmol) in DCE (1.4 mL) reacted at 75 °C for 18h. Subsequent flash chromatography (hexanes /AcOEt 7:3, Rf = 0.25) afforded the title compound as a colorless oil (40 mg, 58 %). 1H NMR (250 MHz, CD3CN, 65 °C) δ 7.42 – 7.20 (m, 5H), 6.85 (d, J = 8.4 Hz, 1H), 5.96 – 5.69 (m, 1H), 5.21 – 5.02 (m, 2H), 4.97 – 4.92 (m, 1H), 4.74 – 4.65 (m, 1H), 4.46 (td, J = 7.4, 3.0 Hz, 1H), 3.96 (dd, J = 5.9, 3.0 Hz, 1H), 3.75 (s, 3H), 3.49 (s, 2H), 2.43 – 2.18 (m, 2H). 13C NMR (62.5 MHz, CD3CN) δ 156.4, 154.6* and 154.3, 136.8, 134.2, 129.6, 129.4, 127.8, 126.7, 118.8, 102.1* and 101.7, 67.0, 66.4, 54.7, 53.9, 38.6, 34.7* and 34.3. [* minor rotamer]. HRMS (ESI) m/z [M + Na+] Calcd for C18H20N2O4Na 351,1321, found 351.1317.

ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Methyl 2-phenylpiperidine-1-carboxylate (2a). A round-bottomed flask was charged with dioxazine derivative 1a (55 mg, 0.15 mmol), Pd/C 10 % (11 mg), ammonium formate (56 mg, 0.90 mmol) and methanol (1.2 mL). The resulting mixture was allowed to stir 1 h at 80 °C. The mixture was taken up with CH2Cl2 and filtered over a short pad of Celite. Removal of solvent afforded a reaction crude that was purified by flash chromatography (hexanes /AcOEt 6:4 + 1% Et3N, Rf = 0.66) to give the title compound (15 mg, 45%). The spectral data are consistent with prior reports.23

(2S*,3R*)-Methyl 2-phenyl-3-((2-phenylacetamido)oxy)piperidine-1-carboxylate (3a). The second eluting fraction of the above flash chromatography (Rf = 0.15) afforded the title compound as a white amorphous solid (22 mg, 39 %). 1H NMR (250 MHz, CDCl3) δ 9.05 (s, 1H), 7.42 – 7.11 (m, 10H), 5.64 (s, 1H), 4.74 (s, 1H), 4.05 – 3.88 (m, 1H), 3.85 – 3.66 (m, 3H), 3.66 – 3.42 (m, 2H), 2.94 – 2.79 (m, 1H), 2.14 –2.00 (m, 1H), 1.96 – 1.57 (m, 2H), 1.34 – 1.17 (m, 1H). 13C NMR (62.5 MHz, CDCl3): δ 168.8, 158.6, 136.8, 134.2, 129.2, 129.0, 127.4, 127.3, 126.5, 126.4, 79.0, 54.5, 53.3, 41.1, 40.4, 23.5, 19.6. HRMS (ESI) m/z [M + Na+] Calcd for C21H24N2O4Na 391.1634; Found 391.1631.

General procedure for the addition of metal hydrides to 1a (Scheme 1) An over-dried 5 mL Schlenk tube was charged with the dioxazine derivative and freshly distilled THF under argon protection. The resulting mixture was cooled at 0 °C and the hydride source was slowly added. Upon disappearance of the dioxazine derivative, the reaction was quenched with water and the aqueous phase was extracted with Et2O. The

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 30

combined organic layers were dried over MgSO4, concentrated and purified by flash chromatography.

(2S*,3R*)-Methyl

2-phenyl-3-((2-phenylacetamido)oxy)-3,4-dihydropyridine-1(2H)-

carboxylate (5a). According to the general procedure, dioxazine 1a (55 mg, 0.15 mmol), LiBH4 (20 mg, 0.9 mmol) and THF (0.26 mL), 0 °C for 18 h. Subsequent flash chromatography (hexanes /AcOEt 6:4, Rf = 0.15) afforded the title compound as an amorphous solid (26 mg, 48 %). 1H NMR (250 MHz, CDCl3) 9.04 (bs, 1H) and 8.60* (bs, 1H) (exchangeable with D2O), 7.39 – 7.21 (m, 8H), 7.18 – 7.08 (m, 2H), 7.02 (d, J = 7.4 Hz, 1H), 5.59 (s, 1H) and 5.44* (s, 1H), 4.77 (bs, 1H), 4.44 (bs, 1H), 3.74 (s, 3H) and 3.61* (s, 3H), 3.60 - 3.48 (m, 2H), 2.38 – 2.11 (m, 1H), 2.08-1.79 (m, 1H).

13C

NMR (62.5 MHz, CDCl3) δ

169.5, 154.7 and 154.4*, 138.4 and 138.2*, 134.2, 129.3, 129.0, 128.9, 127.8, 127.4, 125.7, 125.3, 124.7, 102.6 and 102.0*, 79.9 and 79.4*, 56.4 and 55.8*, 53.5, 40.9, 22.0 and 21.3*. [*Minor rotamer] HRMS (ESI) m/z [M + Na+] Calcd for C21H22N2O4Na 389.1477; Found 389.1472.

General procedure for the synthesis of [3.3.1]-bicyclic derivatives 6a-g (Scheme 2) An over-dried 5 mL Schlenk tube was charged with the dioxazine derivative and freshly distilled THF under argon protection. The resulting mixture was cooled at 0 °C and a solution of LiBHEt3 (1.0 M in THF, 6.0 eq.) was dropwise added (CAUTION PYROFORIC REAGENT). Upon disappearance of the dioxazine derivative, the reaction was quenched with water and the aqueous phase was extracted with Et2O. The combined organic layers were dried over MgSO4, concentrated and purified by flash chromatography.

ACS Paragon Plus Environment

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

2-Phenyl-1-((1R*,5S*,9R*)-9-phenyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3yl)ethanone (6a). According to the general procedure, dioxazine 1a (365 mg, 1.0 mmol), LiBHEt3 (6.0 ml, 6.0 mmol) and THF (1.6 mL), 0°C for 1h. Subsequent flash chromatography (hexanes /AcOEt 8:2, Rf = 0.17) afforded the title compound as a white solid (321 mg, 76 %). M.p. = 156-159 °C. 1H NMR (250 MHz, CDCl3) δ 7.39 – 7.20 (m, 10H), 6.07 – 5.95 (m, 2H), 5.61 (d, 1H, J = 13.2 Hz), 4.99 – 4.89 (m, 1H), 4.57 (d, 1H, J = 2.2 Hz), 4.48 (d, 1H, J = 13.2 Hz), 3.82 (d, 1H, J = 14.7 Hz), 3.55 (d, 1H, J = 14.7 Hz), 3.45 – 3.18 (m, 2H).

13C

NMR

(62.5 MHz, CDCl3) δ 172.3, 137.0, 129.7, 128.7, 128.5, 128.4, 128.0, 127.7, 126.9, 126.8, 119.4, 72.3, 65.7, 59.6, 47.6, 39.8. HRMS (ESI) m/z [M + Na+] Calcd for C20H20N2O2Na 343.1422; Found 343.1421.

2-Phenyl-1-((1R*,5S*,8R*,9R*)-9-phenyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl2,2,8-D3)ethan-1-one (6a-D). According to the general procedure, dioxazine 1a (59 mg, 0.16 mmol), LiBDEt3 (0.97 ml, 0.97 mmol) and THF (0.28 mL), 0°C for 1 h. Subsequent crystallization (hexanes/AcOEt) afforded the title compound as a white solid (11 mg, 21.3%). M.p.=159 °C. 1H NMR (250 MHz, CDCl3) δ 7.41 – 7.20 (m, 10H), 6.07 – 5.94 (m, 2H), 4.96 – 4.88 (m, 1H), 4.59 – 4.52 (m, 1H), 3.82 (d, J = 14.7 Hz, 1H), 3.54 (d, J = 14.7 Hz, 1H), 3.31 (s, 1H).

13C

NMR (62.5 MHz, CDCl3) δ 172.3, 137.1, 137.0, 135.1, 129.7, 128.5, 128.3,

127.6, 126.9, 126.7, 119.4, 72.3, 59.5, 47.25 (J = 21.1 Hz), 39.8. HRMS (ESI) m/z [M + Na+] Calcd for C20H17D3N2O2Na 346.1611; Found 346.1608.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 30

1-((1R*,5S*,9R*)-9-Ethyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl)-2phenylethanone (6b). According to the general procedure, dioxazine 1d (50 mg, 0.15 mmol), LiBHEt3 (0.90 ml, 0.90 mmol) and THF (0.26 mL), 0°C for 30 minutes. Subsequent flash chromatography (hexanes /AcOEt 5:5, Rf = 0.30) afforded the title compound as a white solid (27 mg, 65 %). M.p. = 68-70 °C. 1H NMR (250 MHz, CDCl3) δ 7.45 – 7.10 (m, 5H), 6.20 – 6.13 (m, 1H), 5.91–5.77 (m, 1H), 5.48 (d, 1H, J = 13.1 Hz), 4.33 – 4.19 (m, 2H), 3.77 (d, 1H, J = 14.7 Hz), 3.61 - 3.46 (m, 2H), 3.45 – 3.31 (m, 1H), 3.25 (app-t, 1H, J = 7.4 Hz), 1.511.33 (m, 2H), 0.99 (t, 3H J = 7.4 Hz). 13C NMR (62.5 MHz, CDCl3) δ 172.2, 136.5, 135.1, 129.6, 128.5, 126.8, 119.0, 73.6, 65.9, 59.5, 47.5, 39.6, 22.7, 10.6. HRMS (ESI) m/z [M + Na+] Calcd for C16H20N2O2Na 295.1422; Found 295.1421.

1-((1R*,5S*)-4-Oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl)-2-phenylethanone

(6c).

According to the general procedure, dioxazine 1e (53 mg, 0.15 mmol), LiBHEt3 (0.90 mL, 0.90 mmol) and THF (0.26 mL), 0°C for 30 minutes. Subsequent flash chromatography (hexanes /AcOEt 5:5 + 5% Et3N, Rf = 0.20) afforded the title compound as a white solid (13 mg, 35 %). M.p. = 95-98 °C. 1H NMR (250 MHz, CDCl3) δ 7.39 – 7.18 (m, 5H), 6.22 (dt, 1H J =10.0, 5.4 Hz), 6.03 - 5.91 (m , 1H), 5.39 (d, J = 13.2 Hz, 1H), 4.36 (d, J = 6.1 Hz, 1H), 4.30 (d, J = 13.2 Hz, 1H), 3.83 – 3.42 (m, 5H), 3.0 (d, 1H, J= 13.5 Hz). 13C NMR (62.5 MHz, CDCl3) δ 172.1, 136.8, 135.0, 129.6, 128.5, 126.8, 121.1, 70.1, 64.3, 52.0, 50.0, 39.6. HRMS (ESI) m/z [M + Na+] Calcd for C14H16N2O2Na 267.1109; Found 267.1110

Phenyl((1R*,5S*,9R*)-9-phenyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl)methanone (6d). According to the general procedure, dioxazine 1h (30 mg, 0,08 mmol), LiBHEt3 (0.51

ACS Paragon Plus Environment

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

mL, 0.51 mmol) and THF (0.15 mL), 0 °C for 12 h. Subsequent flash chromatography (hexanes /AcOEt 5:5, Rf = 0.57) afforded the title compound as a white solid (15 mg, 55%). M.p. = 120-122 °C. 1H NMR (250 MHz, CD3CN, 55 °C) δ 7.64 – 7.54 (m, 2H), 7.51 – 7.38 (m, 3H), 7.36 – 7.21 (m, 5H), 5.96 (dt, J = 10.1, 2.7 Hz, 1H), 5.75 – 5.54 (m, 3H), 4.82 (d, J = 6.4 Hz, 1H), 4.69 (d, J = 13.0 Hz, 1H), 4.63 (s, 1H), 3.39 (d, J = 20.5 Hz, 1H), 3.23 (d, J = 19.6 Hz, 1H). 13C NMR (62.5 MHz, CD3CN) δ 172.3, 138.3, 137.1, 136.0, 131.2, 129.9, 129.0, 128.7, 128.2, 127.6, 121.3, 73.1, 66.9, 60.0, 48.4. HRMS (ESI) m/z [M + Na+] Calcd for C19H18N2O2Na 329.1266; Found 329,1261.

2-Phenyl-1-((1R*,5S*,9R*)-9-vinyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl)ethanone (6e). According to the general procedure, dioxazine 1i (110 mg, 0.35 mmol), LiBHEt3 (2.1 ml, 2.1 mmol) and THF (0.6 mL), 0 °C for 7 h. Subsequent flash chromatography (petroleum ether /AcOEt 5:5, Rf = 0.20) afforded the title compound as a colorless oil (38 mg, 40%). 1H NMR (250 MHz, CDCl3) δ 7.36 – 7.17 (m, 5H), 6.18 – 6.08 (m, 1H), 5.94 – 5.83 (m, 1H), 5.69 (ddd, J = 17.4, 10.6, 5.4 Hz, 1H), 5.47 (d, J = 13.3 Hz, 1H), 5.36 – 5.21 (m, 2H), 4.37 – 4.32 (m, 1H), 4.29 (d, J = 13.1 Hz, 1H), 3.96 (dd, J = 3.4, 1.4 Hz, 1H), 3.75 (d, J = 14.8 Hz, 1H), 3.65 – 3.53 (m, 1H), 3.48 (d, J = 14.7 Hz, 1H), 3.44 – 3.31 (m, 1H). 13C NMR (62.5 MHz, CDCl3) δ 172.2, 136.7, 135.0, 133.2, 129.6, 128.5, 126.8, 119.0, 118.7, 72.9, 65.2, 59.5, 47.9, 39.6. HRMS (ESI) m/z [M + Na+] Calcd for C16H18N2O2Na 293.1266; Found 293.1261.

2-Phenyl-1-((1R*,5S*,9R*)-9-cyclohexyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3yl)ethanone (6f). According to the general procedure, dioxazine 1j (155 mg, 0.42 mmol), LiBHEt3 (2.5 ml, 2.5 mmol) and THF (0.72 mL), 0 °C for 6.5 h. Subsequent flash

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

chromatography (petroleum ether /AcOEt 7:3, Rf = 0.17) afforded the title compound as a withe solid (32 mg, 23%; m.p. = 139-142°C) and starting material 1j (Conv: ca 50%, Rf = 0.23). 1H NMR (250 MHz, CDCl3) δ 7.36 – 7.19 (m, 5H), 6.20 – 6.10 (m, 1H), 5.90 – 5.78 (m, 1H), 5.45 (d, J = 13.0 Hz, 1H), 4.41 (d, J = 5.9 Hz, 1H), 4.19 (d, J = 13.1 Hz, 1H), 3.75 (d, J = 14.7 Hz, 1H), 3.55 – 3.26 (m, 3H), 2.95 (d, J = 10.3 Hz, 1H), 1.96 (d, J = 12.5 Hz, 1H), 1.80 – 1.56 (m, 4H), 1.38 – 1.09 (m, 4H), 1.06 – 0.78 (m, 2H). 13C NMR (62.5 MHz, CDCl3) δ 172.1, 137.0, 135.2, 129.6, 128.4, 126.7, 118.9, 72.5, 66.2, 63.0, 47.6, 39.6, 36.3, 30.2, 29.1, 26.5, 25.9, 25.9. HRMS (ESI) m/z [M + Na+] Calcd for C20H26N2O2Na 349.1892; Found 349.1887.

2-Phenyl-1-((1R*,5S*,9R*)-9-allyl-4-oxa-1,3-diazabicyclo[3.3.1]non-6-en-3-yl)ethanone (6g). According to the general procedure, dioxazine 1k (115 mg, 0.35 mmol), LiBHEt3 (2.1 ml, 2.1 mmol) and THF (0.6 mL), 0 °C for 1.5 h. Subsequent flash chromatography (hexane/AcOEt 6:4, Rf = 0.13) afforded the title compound as a colorless oil (72 mg, 72 %). 1H

NMR (250 MHz, CDCl3) δ 7.36 – 7.13 (m, 5H), 6.17 (dt, J = 10.0, 2.6 Hz, 1H), 5.91 – 5.72

(m, 2H), 5.45 (d, J = 13.1 Hz, 1H), 5.18 – 5.04 (m, 2H), 4.30 – 4.15 (m, 2H), 3.74 (d, J = 14.7 Hz, 1H), 3.59 – 3.32 (m, 4H), 2.27 – 2.02 (m, 2H). 13C NMR (62.5 MHz, CDCl3) δ 172.1, 136.5, 135.0, 134.3, 129.6, 128.4, 126.7, 118.7, 117.6, 73.1, 65.8, 57.5, 47.5, 39.6, 34.4. HRMS (ESI) m/z [M + Na+] Calcd for C17H20N2O2Na 307.1422; Found 307.1419.

ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

General procedure for the addition of metal hydrides to 1a in the presence of palladium catalysts (Table 1) An over-dried 5 mL Schlenk tube was charged with the dioxazine derivative, the palladium catalyst (0.05 eq) and freshly distilled THF under argon protection. The resulting mixture was cooled at 0 °C and the hydride source was slowly added. Upon disappearance of the dioxazine derivative, the reaction was quenched with water and the aqueous phase was extracted with Et2O. The combined organic layers were dried over MgSO4, concentrated and purified by flash chromatography.

(5R*,6S*)-Methyl

6-phenyl-5-((2-phenylacetamido)oxy)-5,6-dihydropyridine-1(2H)-

carboxylate (4a). According to the general procedure, dioxazine 1a (54.7 mg, 0.15 mmol), (PPh3)2PdCl2 (5.4 mg, 0.078 mmol), NaBH4 (35 mg, 0.9 mmol) and THF (0.75 mL) for 24 h at rt. Subsequent flash chromatography (hexanes /AcOEt 6:4, Rf = 0.22) afforded the title compound as sticky oil (16 mg, 30%). The slower eluting fractions afforded compound 5a (vide supra). 1H NMR (250 MHz, CD3CN, 65 °C) δ 9.21 (bs, 1H), 7.39 – 7.24 (m, 8H), 7.21 – 7.13 (m, 2H), 6.14* (dd, J = 3.5, 2.7 Hz, 1H) and 6.10 (dd, J = 3.5, 2.7 Hz, 1H), 6.03 – 5.93 (m, 1H), 5.67 (s, 1H), 4.63 (d, J = 5.7 Hz, 1H), 4.42 – 4.37* (m, 1H) and 4.34 – 4.30 (m, 1H), 3.71 (s, 3H), 3.66 (dd, J = 4.0, 1.9 Hz, 1H) and 3.59* (dd, J = 4.0, 2.0 Hz, 1H), 3.48 (s, 2H). 13C NMR (62.5 MHz, CDCl3) δ 169.0, 157.0, 136.7, 134.1, 132.2, 129.2, 128.9, 128.7, 127.9, 127.4, 127.2, 121.1, 53.5, 53.2, 41.25, 31.0, 29.8. HRMS (ESI) m/z [M + Na+] Calcd for C21H22N2O4Na 389.1477; Found 389.1478.

Compound 4a-D. According to the general procedure, dioxazine 1a (57 mg, 0.15 mmol),

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(PPh3)2PdCl2 (5.4 mg, 0.078 mmol), NaBD4 (60 mg, 1.4 mmol) and THF (0.75 mL), rt for 8 h. Subsequent flash chromatography (hexanes /AcOEt 6:4, Rf = 0.19) afforded the title compound as white amorphous solid (8 mg, 15%). 1H NMR (250 MHz, CD3CN, 65 °C) δ 9.22 (s, 1H), 7.38 – 7.24 (m, 8H), 7.21-7.13 (m, 2H), 6.11 (dd, J = 10.2, 2.2 Hz, 1H), 6.03 – 5.93 (m, 1H), 5.68 (s, 1H), 4.64 (dt, J = 5.5, 1.7 Hz, 1H), 3.71 (s, 3H), 3.64-3.56 (m, 1H), 3.48 (s, 2H). 13C

NMR (75 MHz, CD3CN) δ 169.6, 157.5, 139.6, 136.4, 132.7, 130.1, 129.6, 129.5, 128.5,

127.8, 127.5, 121.2, 79.5, 57.5, 55.8, 53.4, 42.1 (J = 21.9 Hz), 40.8. HRMS (ESI) m/z [M + Na+] Calcd for C21H21DN2O4Na 390.1540; Found 390.1538.

Compound 5a-D. From the slower eluting fraction of the above flash chromatography (hexanes /AcOEt 6:4, Rf = 0.10) was recovered the title compound as white amorphous solid (12 mg, 22%). 1H NMR (250 MHz, CDCl3) δ 8.92 (bs, 1H) and 8.50* (bs, 1H) (exchangeable with D2O), 7.40 – 7.26 (m, 8H), 7.18 –7.02 (d, J = 8.4 Hz, 1H), 5.60 (s, 1H) and 5.45* (s, 1H), 4.93 – 4.71 (m, 1H), 4.45 (bs, 1H), 3.74 (s, 3H) and 3.61* (s, 3H), 3.52 (d, J = 18.3 Hz, 2H), 2.21 (d, J = 15.2 Hz, 1H). 13C NMR (75 MHz, CDCl3) δ 169.6, 154.7* and 154.5, 138.4* and 138.1, 134.1, 129.3, 129.1, 128.9, 127.8, 127.6, 125.7, 125.4, 102.5* and 101.9, 79.8, 56.5* and 55.8, 53.5, 41.1, 22.1-21.1 (m). [* minor rotamer] HRMS (ESI) m/z [M + Na+] Calcd for C21H21DN2O4Na 390.1540; Found 390.1538.

Supporting Information Available: The Supporting Information is available free of charge on the ACS Publications website at DOI: Computational and X-ray data, copies of 1H and 13C NMR spectra of all new compounds, and 1H NMR spectrum of known compound 2a.

■ AUTHOR INFORMATION

ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Corresponding Author *E-mail: [email protected] ORCID Mauro Pineschi: orcid.org/0000-0001-6063-2198

Notes The authors declare no competing financial interest.

Acknowledgement. We gratefully acknowledge the financial support from P.R.A. 2016_27 by the University of Pisa. We also gratefully acknowledge Eli Lilly Company for providing us biological data on GLP-1 secretion within the Open Innovation Drug Discovery Program used with Lilly's permission. (https://openinnovation.lilly.com/dd/).

REFERENCES 1) Vitaku, E.; Smith, D. T.; Njardarson, J. T. Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem. 2014, 57, 10257-10274. 2) Selected recent papers: (a) Beng, T. K.; Takeuchi, H.; Weber, M.; Sarpong, R. Stereocontrolled Synthesis of Vicinally Functionalized Piperidines by Nucleophilic βAddition of Alkyllithiums to α-Aryl Substituted Piperidine Enecarbamates. Chem. Commun. 2015, 51, 7653-7656. (b) Parikh, N.; Sudipta Raha, R.; Kapileswar, S.; Chakraborti, A. K. Multicomponent Diastereoselective Synthesis of Tetrahydropyridines in Aqueous Micelles. Synthesis 2016, 48, 547-548. (c) Berti, F.; Malossi, F.; Marchetti, F.; Pineschi, M. A highly Enantioselective Mannich Reaction of Aldehydes with Cyclic N-Acyliminium Ions by

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Synergistic Catalysis. Chem. Commun. 2015, 51, 13694-13697. (d) Barange, D. K.; Johnson, M. T.; Cairns, A. G.; Olsson, R.; Almqvist, F. Regio- and Stereoselective Alkylation of PyridineN-oxides: Synthesis of Substituted Piperidines and Pyridines. Org. Lett. 2016, 18, 62286231. (e) Tait, M. B.; Butterworth, S.; Clayden, J. 2,2- and 2,6-Diarylpiperidines by Aryl Migration within Lithiated Urea Derivatives of Tetrahydropyridines. Org. Lett. 2015, 17, 1236-1239. (f) Chen, S.; Mercado, B. Q.; Bergman, R. G.; Ellman, J. A. Regio- and Diastereoselective Synthesis of Highly Substituted, Oxygenated Piperidines from Tetrahydropyridines. J. Org. Chem. 2015, 80, 6660-6668. (g) Lei, C.-H.; Wang, D.-X.; Zhao, L.; Zhu, J.; Wang, M.-X. Synthesis of Multifunctionalized 1,2,3,4-Tetrahydropyridines, 2,3Dihydropyridin-4(1H)-ones, and Pyridines from Tandem Reactions Initiated by [5+1] Cycloaddition of N-Formylmethyl-Substituted Enamides to Isocyanides: Mechanistic Insight and Synthetic Application. Chem. Eur. J. 2013, 19, 16981-16988. 3) (a) Crotti, S.; Berti, F.; Pineschi, M. Copper-catalyzed Perkin-acyl-Mannich Reaction of Acetic Anhydride with pyridine: Expeditious Entry to Unconventional Piperidines. Org. Lett. 2011, 13, 5152-5155. (b) Berti, F.; Di Bussolo, V.; Pineschi, M. Synthesis of Protected (1Phenyl-1H-pyrrol-2-yl)-alkane-1-amines from Phenylnitroso Diels–Alder Adducts with 1,2Dihydropyridines. J. Org. Chem. 2013, 78, 7324-7329. (c) Berti, F.; Di Bussolo, V.; Pineschi, M. Synthesis of 2,7-Diazabicyclo[2.2.1]heptenes by N–O Bond Cleavage of Arylnitroso Diels–Alder 1,2-Dihydropyridine Cycloadducts. Synthesis 2015, 47, 647-652. 4) For a recent catalytic enantioselective C2-arylation, see: Lutz, P.; Chau, S. T.; Doyle, A. G. Nickel-Catalyzed Enantioselective Arylation of Pyridine. Chemical Science 2016, 7, 41054109.

ACS Paragon Plus Environment

Page 26 of 30

Page 27 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

5) Bull, J. A.; Mousseau, J. J.; Pelletier, G.; Charette, A. B. Synthesis of Pyridine and Dihydropyridine Derivatives by Regio- and Stereoselective Addition to N-Activated Pyridines. Chem. Rev. 2012, 112, 2642-2713. 6) (a) Dubey, S. K.; Knaus, E. E. Some Regio- and Stereochemical Aspects of the Diels-Alder Reaction of Nitrosocarbonyl Compounds with N-Substituted 1,2-Dihydropyridines. J. Org. Chem. 1985, 50, 2080-2086. (b) Backenstrass, F.; Streith, J.; Tschamber, T. Stereospecific Double Glycolisation of 1,2-Dihydropyridines with OsO4. Synthesis of (±) Aminoarabinoseand of (±) Aminoaltrose Derivatives. Tetrahedron Lett. 1990, 31, 2139-2142. 7) Entwistle, I. D.; Wood, W. W. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, 1991; Vol. 8, pp 955- 981 and references therein. 8) For palladium-catalyzed reduction of cyclic systems, see: (a) Greenspoon, N.; Keinan, E. Selective Deoxygenation of Unsaturated Carbohydrates with Pd(0)/Ph2SiH2/ZnCl2. Total Synthesis of (+)-(S,S)-(6-Methyltetrahydropyran-2-yl)acetic acid. J. Org. Chem 1988, 53, 3723-3731. (b) Hanessian, S.; Maianti, J. P.; Matia, R. D.; Feeney, L. A.; Armstrong, E. S. Hybrid Aminoglycoside Antibiotics via Tsuji Palladium-Catalyzed Allylic Deoxygenation. Org. Lett. 2011, 13, 6476-6479. 9) Caspi, D. D.; Garg, N. K.; Stoltz, B. M. Heterogeneous Reductive Isomerization Reaction Using Catalytic Pd/C and H2. Org. Lett. 2005, 7, 2513-2516. 10) Berti, F.; Menichetti, A.; Di Bussolo, V.; Favero, L.; Pineschi, M. Synthesis of Bicyclic Piperidinyl Enamides and Enecarbamates by Hetero-Cope Rearrangement of Nitroso Cycloadducts. Chem. Heterocyc. Compd. 2018, 54, 458-468 and references therein.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 30

11) For a seminal regioselective palladium-catalyzed hydrogenolysis of allylic formate, see: Mandai, T.; Matsumoto, T.; Kawada, M.; Tsuji, J. Stereocontrolled Formation of cis and trans Ring Junctions in Hydrindane and Decalin Systems by Palladium-Catalyzed Regioselective and Stereospecific Hydrogenolysis of Allylic Formates. J. Org. Chem. 1992, 57, 1326. 12) (a) Kim, H. J.; Su, L.; Jung, H.; Koo, S. Selective Deoxygenation of Allylic Alcohol: Stereocontrolled Synthesis of Lavandulol. Org. Lett. 2011, 13, 2682-2685. (b) Wipf, P.; Spencer,

S.

R.

Asymmetric

Total

Syntheses

of

Tuberostemonine,

Didehydrotuberostemonine, and 13-Epituberostemonine. J. Am. Chem. Soc. 2005, 127, 225235. 13) See for examples: (a) Huang, W.-X.; Wu, B.; Gao, X.; Chen, M.-W.; Wang, B.; Zhou, Y.-G. Iridium-Catalyzed Selective Hydrogenation of 3-Hydroxypyridinium Salts: A Facile Synthesis of Piperidin-3-ones. Org. Lett. 2015, 17, 1640-1643. (b) Hardy, S.; Martin, S. F. Multicomponent, Mannich-type assembly process for generating novel, biologically-active 2-arylpiperidines and derivatives. Tetrahedron 2014, 70, 7142-7157. (c) Huy, P. H.; Westphal, J. C.; Koskinen, A. M. P. Concise, Stereodivergent and Highly Stereoselective Synthesis of cis- and trans-2-Substituted 3-Hydroxypiperidines – Development of a Phosphite-Driven Cyclodehydration. Beilstein J. Org. Chem. 2014, 10, 369-383. 14) CCDC 1587512 contains the supplementary crystallographic data for this paper. Crystal data for 6a: C20H20N2O2, Mr = 320.38, monoclinic, space group P21/c (no. 14), a = 10.9704(3) Å, b = 6.5342(2) Å, c = 23.6337(7) Å, β = 101.584(2), V = 1659.62(8) Å3, Z = 4, F(000) = 680, Dc = 1.282 g⋅cm−3, µ (Mo-Kα) = 0.084 mm−1, T = 296 K, 19785 reflections

ACS Paragon Plus Environment

Page 29 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

collected, 5574 unique (Rint = 0.0197), 232 variables refined with 3963 reflections with Fo > 4σ(Fo) to R1 = 0.0451. CCDC 1587512 for more details. 15) Relative stereochemistry determined by 2D NOESY/ROESY experiments. 16) The hydrogen of B-H bond has an increased hydride character in Super-H® than in BH4-, and it is also more tightly coordinated to the cation. For example, see : Golub, I. E.; Filippov, O. A.; Gulyaeva, E.; Gutsul, E. I.; Belkova, N. V. The Interplay of Proton Accepting and Hydride Donor Abilities in the Mechanism of Step-wise Boron Hydrides Alcoholysis. Inorganica Chimica Acta 2017, 456, 113-119 and references therein. 17) See Supporting Information for details. 18) Brown, H. C.; Kim, S. C.; Krishnamurthy, S. J. Selective reductions. 26. Lithium Triethylborohydride as an Exceptionally Powerful and Selective Reducing Agent in Organic Synthesis. Exploration of the Reactions with Selected Organic Compounds Containing Representative Functional Groups. J. Org. Chem. 1980, 45, 1-12. 19) Arase, A.; Hoshi, M.; Yamaki, T.; Nakanishi, H. J. Lithium Borohydride-Catalysed Selective Reduction of Carbonyl Group of Conjugated and Unconjugated Alkenones with Borane in Tetrahydrofuran. Chem. Soc., Chem. Commun. 1994, 855-856. 20) Brown, H. C.; Krishnamurthy, S.; Hubbard, J. L. Addition Compounds of Alkali Metal Hydrides. 15. Steric Effects in the Reaction of Representative Trialkylboranes with Lithium and Sodium Hydrides to Form the Corresponding Trialkylborohydrides. J. Am. Chem. Soc. 1978, 100, 3343-3349.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

21) Yamaguchi, R.; Nakazono, Y.; Matsuki, T.; Hata, E.; Kawanisi, M. Highly Regioselective α-Addition of Alkynyl and Alkenyl Grignard Reagents to 1-Alkoxycarbonylpyridinium Salts and Its Application to Synthesis of 1-Azabicycloalkanes and (±)-Solenopsin A. Bull. Chem. Soc. Jap. 1987, 60, 215-222.

22) Yamaguchi, R.; Moriyasu, M.; Yoshioka, M.; Kawanisi, M. Reaction of Allylic Tin Reagents with Nitrogen Heteroaromatics Activated by Alkyl Chloroformates: Regioselective Synthesis of α-Allylated 1,2-Dihydropyridines and Change of the Regioselectivity Depending on Methyl Substituents at the Allylic Moiety. J. Org. Chem. 1988, 53, 3507-3512.

23) Suga, S.; Okajima, M.; Yoshida, J. Reaction of an Electrogenerated “Iminium Cation Pool” with Organometallic Reagents. Direct Oxidative α-Alkylation and –Arylation of Amine Derivatives. Tetrahedron Lett. 2001, 42, 2173-2176.

ACS Paragon Plus Environment

Page 30 of 30